PCGEM1, a prostate-specific gene, is overexpressed in prostate cancer

PCGEM1, a prostate-specific gene, is overexpressed in
prostate cancer
Vasantha Srikantan*†, Zhiqiang Zou*†, Gyorgy Petrovics*, Linda Xu*, Meena Augustus*‡, Leland Davis*,
Jeffrey R. Livezey*, Theresa Connell*, Isabell A. Sesterhenn§, Kiyoshi Yoshino¶, Gregory S. Buzard¶,
F. K. Mostofi§, David G. McLeod*㥋, Judd W. Moul*㥋, and Shiv Srivastava*,**
*Center for Prostate Disease Research, Department of Surgery, Uniformed Services University of the Health Sciences, Bethesda, MD 20814-4799; ‡Genetics
Department, Division of Clinical Sciences, National Cancer Institute, National Institutes of Health, Bethesda, MD 20892-4470; §Armed Forces Institute of
Pathology, Department of Genitourinary Pathology, Washington, DC 20307; ¶Laboratory of Comparative Carcinogenesis, Science Applications
International Corporation, National Cancer Institute, Frederick Cancer Research and Development Center, Frederick, MD 21702-1201; and
㛳Urology Service, Department of Surgery, Walter Reed Army Medical Center, Washington, DC 20307-5001
Communicated by Bernhard Witkop, National Institutes of Health, Chevy Chase, MD, August 17, 2000 (received for review February 28, 2000)
A prostate-specific gene, PCGEM1, was identified by differential
display analysis of paired normal and prostate cancer tissues.
Multiple tissue Northern blot analysis revealed that PCGEM1 was
expressed exclusively in human prostate tissue. Analysis of PCGEM1 expression in matched normal and primary tumor specimens
revealed tumor-associated overexpression in 84% of patients with
prostate cancer by in situ hybridization assay and in 56% of
patients by reverse transcription–PCR assay. Among various prostate cancer cell lines analyzed, PCGEM1 expression was detected
only in the androgen receptor-positive cell line LNCaP. Extensive
DNA sequence analysis of the PCGEM1 cDNA and genomic DNA
revealed that PCGEM1 lacks protein-coding capacity and suggests
that it may belong to an emerging class of noncoding RNAs, also
called ‘‘riboregulators.’’ The PCGEM1 locus was mapped to chromosome 2q32. Taken together, the remarkable prostate-tissue
specificity and androgen-dependent expression of PCGEM1 as well
as its elevated expression in a significant percentage of tumor
tissues suggest specific functions of PCGEM1 in the biology and
tumorigenesis of the prostate gland.
riboregulator 兩 differential display 兩 androgen regulation 兩 noncoding RNA
P
rostate cancer (CaP) is the most common malignancy in men
in the United States and the second leading cause of cancer
mortality (1). The serum prostate-specific antigen (PSA) test has
revolutionized the early detection of CaP (2). Because of the
prostatic epithelial-cell-specific expression of PSA, it is also of
value as a biomarker for disease remission兾progression after
treatment (2). Although PSA is effective in identifying men who
may have CaP, it is often elevated in men with benign prostatic
hyperplasia, prostatitis, and other nonmalignant disorders (2).
Therefore, identification of additional CaP-specific molecular
markers is needed to refine the diagnosis as well as prognosis for
CaP. It is also recognized that early detection of CaP presents
challenges with respect to predicting the clinical course of
disease for individual patients (2). The wide spectrum of biologic
behavior exhibited by prostatic neoplasms calls for the identification of biomarkers that may be able to distinguish a slow
growing cancer from a more aggressive cancer with a potential
to metastasize (2). CaP-associated molecular genetic alterations
are being unraveled by using various strategies involving: (i)
analyses of genes commonly involved in human cancer, (ii)
positional cloning of putative genes on frequently affected
chromosome loci in CaP, and (iii) gene expression profiling of
normal and tumor specimens of patients with CaP (3).
The discovery of additional prostate-specific genes has also
resulted in enthusiasm for evaluating their potential in CaP
diagnosis and disease progression. Increased expression of prostate-specific membrane antigen (PSMA) has been correlated
with more aggressive CaP (4). CaP-associated expression of
PSMA is being evaluated for imaging of CaP by radiolabeled
12216 –12221 兩 PNAS 兩 October 24, 2000 兩 vol. 97 兩 no. 22
anti-PSMA monoclonal antibodies (4). Furthermore, promising
immunotherapy approaches are being pursued with PSMA
peptides (5). Recently described prostate-specific genes include
an androgen-regulated homeobox gene, NKX3.1, which exhibits
function(s) in mouse prostate growth兾development (6, 7) and
shows overexpression in a subset of CaP (26). A prostate-specific
gene, prostate stem cell antigen, has been shown to exhibit
overexpression in CaP (8). The prostate-specific兾androgenregulated serine proteases, prostase and TMPRSS2 (9, 10), and
a prostate-specific gene, DD3 (11), represent the latest additions
to a small number of reports describing the discovery of prostatespecific genes.
In this report, we describe the discovery of a prostate-specific
gene, PCGEM1, which was identified during our characterization of CaP-associated gene expression alterations. The most
striking aspect of PCGEM1 characterization represents its prostate-tissue-specific expression with restricted expression in glandular epithelial cells. Computational analysis of PCGEM1 cDNA
sequence revealed unexpected characteristics that suggest that
PCGEM1 does not code for a protein, a property similar to a
recently described prostate-specific gene, DD3 (11). Thus, PCGEM1 and DD3 may define a previously uncharacterized class
of prostate-tissue-specific genes whose functions remain to be
determined. Elevated PCGEM1 expression in a significant fraction of CaP specimens further suggests the role of PCGEM1
overexpression in prostate epithelial cell proliferation and兾or
tumorigenesis.
Materials and Methods
CaP Specimens and Human CaP Cell Lines. Matched CaP and adja-
cent normal prostate tissues were obtained from patients who
had undergone radical prostatectomy at Walter Reed Army
Medical Center. Tissue histopathology and microdissections
were performed as described (12). LNCaP, DU145, and PC-3
CaP cell lines were obtained from American Type Culture
Collection. DuPro-1 is a nude mice xenograft (13). CPDR-1 is a
primary CaP-derived cell line immortalized by retroviral vector
LXSN 16 E6 E7 expressing the E6 and E7 genes of the human
papilloma virus 16 (a gift from D. Galloway, Fred Hutchinson
Abbreviations: CaP, prostate cancer; PSA, prostate-specific antigen; PSMA, prostate-specific membrane antigen; DD, differential display; RT-PCR, reverse transcription–PCR;
GAPDH, glyceraldehyde-3-phosphate dehydrogenase.
Data deposition: The sequence reported in this paper has been deposited in the GenBank
database (accession no. AF223389).
†V.S.
and Z.Z. contributed equally to this work.
**To whom reprint requests should be addressed. E-mail: [email protected].
The publication costs of this article were defrayed in part by page charge payment. This
article must therefore be hereby marked “advertisement” in accordance with 18 U.S.C.
§1734 solely to indicate this fact.
Preparation of RNA and Differential Display (DD) Analysis. Total
RNA was prepared from Optimal Cutting Temperature Compound (Miles, Diagnostics Division, Elkhart, IN) embedded
frozen tissues for the DD analysis. Histologically defined
matched normal and tumor tissues were initially quantified for
the presence of epithelial cells by using hematoxylin- and
eosin-stained slides. Genomic DNA-free total RNA was isolated
from the enriched pool of cells derived from normal and tumor
tissue sections, and the epithelial nature of the RNA source was
confirmed further by using human cytokeratin-18 expression in
reverse transcription–PCR (RT-PCR) assays. To analyze PCGEM1 expression in CaP specimens, hematoxylin- and eosinstained slides corresponding to frozen tissue sections were used
as a template, and subsequent unstained sections were superimposed on it. The microscopically defined area of the tumor
cells was excised with a razor blade and used for RNA extraction
by using the RNAzol B method (Tel-Test, Friendswood, TX).
Poly(A)⫹ RNA from different cell lines was prepared by a Fast
Track kit (Invitrogen). DD analysis (14) was performed with the
use of a Hieroglyph mRNA profile kit (Genomyx, Foster City,
CA).
Isolation of Full-Length PCGEM1 cDNA and Sequence Determination.
The full-length cDNA sequence was assembled by 5⬘ and 3⬘ rapid
amplification of cDNA ends method with a Marathon-ready
cDNA kit (CLONTECH). LASERGENE and MACVECTOR DNA
analysis softwares were used to analyze DNA sequences and to
define ORF regions. To obtain full-length cDNA clones, a
normal prostate cDNA library (CLONTECH) was screened
with a 530-bp ␣32P-dCTP-labeled PCGEM1 cDNA fragment,
and the positive clones were characterized by DNA sequencing
with an automated Applied Biosystems 310 sequencer and a
dRhodamine cycle sequencing kit (PE-Applied Biosystems). In
addition to sequence determination of positive clones from the
library, the PCGEM1 cDNA sequence was also determined by
using cDNA from the LNCaP cell line.
Chromosomal Mapping of PCGEM1 Gene by Fluorescence in Situ
Hybridization. A bacterial artificial chromosome clone containing
the PCGEM1 genomic sequence was isolated (Genome Systems, St.
Louis). PCGEM1 bacterial artificial chromosome clone DNA was
nick translated by using spectrum orange (Vysis, Downers Grove,
IL) as a direct label, and fluorescence in situ hybridization was
performed with this probe on normal human male metaphase
chromosome spreads (15). 4⬘,6-Diamidino-2-phenylindole counterstaining was carried out, and chromosomal localization was determined based on the G band analysis of inverted 4⬘,6-diamidino-2phenylindole images (16). NU 200 image-acquisition and registration
software was used to create the digital images. More than 20
metaphases were analyzed.
Analysis of Tissue-Specific Expression of PCGEM1 with Multiple Tissue
Northern Blots. Multiple tissue Northern blots (CLONTECH)
were hybridized with the ␣32P-dCTP-labeled 530-bp PCGEM1
cDNA fragment and other known prostate-specific genes, which
include PSA (77-mer oligo probe), PSMA (234-bp fragment
from PCR product), and NKX3.1 (210-bp cDNA). As an internal
control, glyceraldehyde-3-phosphate dehydrogenase (GAPDH)
probe (17) was also used to hybridize all of the blots.
Analysis of PCGEM1 Expression in Primary CaP and CaP Cell Lines.
DNase-treated RNAs from normal and CaP tissues were used for
the RT-PCR assays. The PCR conditions were optimized to be
within the logarithmic phase of amplification for all of the primers
used. Epithelial-cell-specific cytokeratin-18 expression and
Srikantan et al.
GAPDH expression were used as internal controls. PCR was
performed with Amplitaq Gold from Perkin–Elmer. PCR cycles
were 95°C for 10 min, 1 cycle; 95°C for 30 s, 55°C for 30 s, 72°C for
30 s, 42 cycles; and 72°C for 5 min, 1 cycle followed by 4°C soak
cycle. PCGEM1 PCR primers were 5⬘- TGCCTCAGCCTCCCAAGTAAC-3⬘ (sense) and 5⬘-GGCCAAAATAAAACCAAACAT-3⬘ (antisense). Cytokeratin-18 PCR primers were 5⬘-AGCGCCAGGCCCAGGAGTATGAGG-3⬘ (sense) and 5⬘TATCCGGCGGGTGGTGGTCTTTTG-3⬘ (antisense), and
GAPDH PCR primers were 5⬘-GGGGAGCCAAAAGGGTCATCATCT-3⬘ (sense) and 5⬘-GAGGGGCCATCCACAGTCTTCT-3⬘ (antisense). Cytokeratin-18 expression and PCGEM1 expression were analyzed at 35 and 42 cycles, respectively. PCGEM1
expression was evaluated in different CaP cell lines by RT-PCR as
described above and by Northern blot hybridization with poly(A)⫹
RNA (17).
In Situ Hybridization. In situ hybridization was performed essentially as described by Wilkinson and Green (18). Briefly, Optimal
Cutting Temperature Compound-embedded tissue slides stored
at ⫺80°C were fixed in 4% (vol兾vol) paraformaldehyde, digested
with proteinase K, and then fixed again in 4% (vol兾vol) paraformaldehyde. After washing in PBS, sections were treated with
0.25% acetic anhydride in 0.1 M triethanolamine, washed again
in PBS, and dehydrated in a graded ethanol series. Sections were
hybridized with 35S-labeled riboprobes at 52°C overnight. After
washing and RNase A treatment, sections were dehydrated,
dipped into NTB-2 emulsion, and exposed for 11 days at 4°C.
After development, slides were lightly stained with hematoxylin
and mounted for microscopy. In each section, PCGEM1 expression was scored as percentage of cells showing 35S signal: 1⫹,
1–25%; 2⫹, 25–50%; 3⫹, 50–75%, or 4⫹, 75–100% (an example
of a 2⫹ score is shown in Fig. 6B).
Androgen Regulation of PCGEM1. LNCaP cells were cultured in
RPMI medium 1640 containing 10% (vol兾vol) charcoalstripped FBS for 4 days followed by treatment with R1881, a
nonmetabolizable androgen analog (DuPont) for 12 h and 24 h
at 0.1 nM and 10 nM concentrations. Poly(A)⫹ RNA was
prepared from R1881-treated and untreated cells and was used
in RT-PCR and Northern blot assays.
Results
Isolation and Characterization of the PCGEM1 cDNA. PCGEM1 was
isolated from a DD analysis of paired normal and tumor tissues
of CaP patients. The full length of PCGEM1 cDNA (accession
no. AF223389) was obtained by 5⬘ and 3⬘ rapid amplification of
cDNA ends兾PCR from the original 530-bp DD product by using
cDNA from normal prostate. The rapid amplification of cDNA
ends兾PCR products were sequenced directly. We also used the
530-bp DD fragment as a probe to screen a normal prostate
cDNA library. Three overlapping cDNA clones were identified.
The longest cDNA clone was 1,643 nucleotides in length with a
potential polyadenylation site, ATTAAA, close to the 3⬘ end
followed by a poly(A) tail. A full-length genomic clone was
isolated and sequenced (Z.Z., unpublished work). Comparison
of the cDNA and genomic sequences revealed the organization
of the PCGEM1 transcription unit from three exons (Fig. 1). The
GenBank database searches with BLAST programs did not reveal
any significant homology of PCGEM1 cDNA to a previously
defined gene or protein. A recent search of the high-throughput
genome sequence database revealed perfect homology of PCGEM1 to a chromosome-2-derived, uncharacterized, unfinished
genomic sequence (accession no. AC013401). DNA sequence
analysis of the PCGEM1 cDNA (Fig. 1) and PCGEM1 genomic
DNA (Z.Z., unpublished work) did not reveal a significant long
ORF in either strand. The longest ORF in the sense strand was
PNAS 兩 October 24, 2000 兩 vol. 97 兩 no. 22 兩 12217
MEDICAL SCIENCES
Cancer Research Center, Seattle; L.D. and S.S., unpublished
work).
Fig. 1. Structure of the PCGEM1 transcription unit. Sequence comparison of
the isolated cDNA and genomic DNA clones revealed that the PCGEM1 gene
consists of three exons. kb, kilobase; E, exon; B, BamHI; H, HindIII; X, XbaI; R,
EcoRI.
105 nucleotides (572–679) encoding a 35-aa peptide. However,
the ATG was not in a strong context of initiation.
Evaluation of the Coding Capacity of PCGEM1. The TESTCODE pro-
gram (GCG) identifies potential protein coding sequences of
longer than 200 bases by measuring the nonrandomness of the
composition at every third base, independently from the reading
frames. Analysis of the PCGEM1 cDNA sequence revealed that,
at greater than 95% confidence level, the sequence does not
contain any region with protein-coding capacity (Fig. 2A).
Similar results were obtained when various published noncoding
RNA sequences were analyzed with the TESTCODE program
(data not shown), whereas known protein coding regions of
similar size—i.e., alpha actin (Fig. 2B)—can be detected with
high fidelity. The CODON PREFERENCE program (GCG), which
locates protein-coding regions in a reading-frame-specific manner, further suggested the absence of protein-coding capacity in
the PCGEM1 gene (see www.cpdr.org). In vitro transcription兾translation of PCGEM1 cDNA did not produce a detectable protein兾peptide. Although we cannot unequivocally rule
out whether PCGEM1 codes for a short unstable peptide, at this
time, both experimental and computational approaches strongly
suggest that PCGEM1 cDNA does not have protein-coding
capacity.
Chromosomal Mapping of PCGEM1. To determine the chromosomal
location of the PCGEM1 gene, we used a 200-kilobase bacterial
artificial chromosome clone containing PCGEM1 gene as a
probe for fluorescence in situ hybridization analysis. Consistent
doublet signal was observed on both homologs of chromosome
2 in every cell. Further Giemsa (G) banding pattern confirmed
that PCGEM1 gene was mapped to 2q32 (Fig. 3 A and B). These
results are supported further by the perfect homology of the
PCGEM1 cDNA sequence to a recently described chromosome2-derived genomic sequence (accession no. AC013401).
PCGEM1 Expression is Prostate-Tissue Specific. The distribution of
PCGEM1 mRNA in normal human tissues was examined by
Northern blot analysis. Of the 24 different human tissue mRNA
analyzed (heart, brain, placenta, lung, liver, skeletal muscle,
kidney, pancreas, spleen, thymus, prostate, testis, ovary, small
intestine, colon, peripheral blood, stomach, thyroid, spinal cord,
lymph node, trachea, adrenal gland, and bone marrow), a
1.7-kilobase mRNA transcript specifically hybridizing to the
PCGEM1 cDNA was detected only in prostate tissue (Fig. 4).
Two independent experiments revealed identical results. Further
analysis of RNA Master blot (CLONTECH) containing mRNA
from 50 different tissues from human adult and fetal tissues,
yeast RNA, and Escherichia coli RNA confirmed the prostatetissue specificity of PCGEM1 gene (see www.cpdr.org). Northern
12218 兩 www.pnas.org
Fig. 2. Evaluation of the coding capacity of PCGEM1. Evaluation of the
coding capacity of the PCGEM1 (A) and the human alpha actin (B), independently from the reading frame, by using the TESTCODE program (GCG). The
number of base pairs is indicated on the x axis, the TESTCODE values are shown
on the y axis. Regions of longer than 200 base pairs above the upper line (at
9.5 value) are considered coding; those under the lower line (at 7.3 value) are
considered noncoding at a confidence level greater than 95%.
blot analysis revealed that the prostate-tissue specificity of
PCGEM1 was comparable to the well known prostate marker
PSA and superior to other prostate-specific genes: NKX3.1 (Fig.
4), PSMA (data not shown), prostate stem-cell antigen (8), and
prostase (9). However, PCGEM1 RNA level in prostate tissue
was significantly less in comparison to PSA, PSMA, and NKX3.1.
Evaluation of PCGEM1 Expression in CaP Cell Lines and in Primary CaP.
PCGEM1 gene expression was evaluated in established CaP cell
lines including LNCaP, DU145, PC3, DuPro1, and CPDR-1. In
RT-PCR assays, the PCGEM1 expression was easily detectable
in the widely studied androgen-responsive CaP cell line LNCaP
(see www.cpdr.org). However, PCGEM1 expression was not
detected in CaP cell lines DU145, PC3, and DuPro-1, all of which
lack detectable levels of the androgen receptor.
RT-PCR analysis of microdissected matched normal and
tumor-tissue-derived RNAs from 23 patients with CaP revealed
tumor-associated overexpression of PCGEM1 in 13 (56%) of the
patients (Fig. 5). Of 23 patients, 6 (26%) did not exhibit
detectable PCGEM1 expression in either normal or tumortissue-derived RNAs. Of 23 tumor specimens, 3 (13%) showed
reduced expression in tumors. One of the patients did not exhibit
any change. Expression of housekeeping gene human cytokeratin-18 or GAPDH remained constant in tumor and normal
specimens of all of the patients (Fig. 5). These results were
Srikantan et al.
Fig. 3. Chromosomal localization of PCGEM1. PCGEM1 was mapped by fluorescence in situ hybridization to 2q32 by using a bacterial artificial chromosome
clone. (A) A representative metaphase showing doublet signal on both homologs of chromosome 2 (arrows). (B) A 4⬘,6-diamidino-2-phenylindole-counterstained
chromosome 2 showing the signal (Left). An inverted 4⬘,6-diamidino-2-phenylindole-stained chromosome 2 shown as G bands (Center). An ideogram of
chromosome 2 showing the localization of the signal to band 2q32 (Right).
confirmed further by another set of PCGEM1-specific primers
(data not shown). Importantly, RT-PCR analysis of a panel of
normal and tumor tissues confirmed tumor-associated PCGEM1
overexpression originally observed in the DD experiment.
As a complementary approach, in situ hybridization was
performed to analyze PCGEM1 expression in tissue samples.
Paired normal (benign) and tumor specimens from 13 additional
patients were tested (representative example in Fig. 6). In 11
cases (84%), tumor-associated elevation of PCGEM1 expression
was detected. In 5 of these 11 patients, the expression of
PCGEM1 increased to 1⫹ in the tumor area from an essentially
undetectable level in the normal area, on a 0 to 4⫹ scale (see
Materials and Methods). Tumor specimens from 4 of 11 patients
scored between 2⫹ (as shown in Fig. 6) and 4⫹. Of 11 patients,
2 showed focal signals with a 3⫹ score in the tumor area, and 1
of these patients had similar focal signal (2⫹) in an area
pathologically designated as benign. In the remaining 2 of the 13
cases, there was no detectable signal in any of the tissue areas
tested. Representative in situ hybridization photographs are
available on our web site (www.cpdr.org). PCGEM1 expression
seems to be restricted to glandular epithelial cells in normal and
tumor specimens, and future analysis by in situ methods of higher
resolution will define cell-type specificity of the PCGEM1
expression in prostate gland.
Fig. 4. Tissue-specific expression of PCGEM1. Multiple tissue Northern blots
(CLONTECH) were hybridized with PCGEM1, NKX3.1, and GAPDH cDNA probe.
Blots were exposed to x-ray films for different times: PCGEM1 for 48 h, NKX3.1
for 24 h, and GAPDH for 15 min.
Fig. 5. Evaluation of PCGEM1 expression in primary CaP. Genomic DNA-free
RNA (100-ng) samples from microdissected tissues were used to analyze
expression of PCGEM1 and cytokeratin-18 by RT-PCR. Three independent
experiments showed the same results.
Srikantan et al.
PNAS 兩 October 24, 2000 兩 vol. 97 兩 no. 22 兩 12219
MEDICAL SCIENCES
Regulation of PCGEM1 Expression by Androgens. Prostate-tissue
specificity of PCGEM1 expression prompted us to evaluate
further whether PCGEM1 expression was regulated by androgen, the hormone that plays a critical role in prostate growth and
differentiation. Northern blot analysis of LNCaP cells treated for
24 h with 10 nM synthetic androgen, R1881, showed significant
induction of PCGEM1 expression compared with untreated
samples (Fig. 7). PCGEM1 expression was also evaluated by
Fig. 6. PCGEM1 expression in normal and tumor areas of CaP tissues. In situ hybridization of 35S-labeled PCGEM1 riboprobe to matched normal (A) versus tumor
(B) sections of patients with CaP. The purple areas are hematoxylin-stained cell bodies; the black dots represent the PCGEM1 expression signal. The signal is
background level in the normal (A) and 2⫹ level in the tumor (B) section. The magnification is ⫻40.
RT-PCR after 0.1 nM and 10 nM R1881 treatment. R1881
induced PCGEM1 expression in a dose-dependent manner (see
www.cpdr.org).
Discussion
Herein, we report the identification of an androgen-regulated
prostate-specific gene, PCGEM1, which exhibits overexpression
in a significant percentage of primary CaP specimens. The
striking prostate-tissue specificity of PCGEM1 expression parallels the tissue specificity of PSA and has provided an impetus
for a comprehensive molecular characterization of PCGEM1.
Prostate-tissue-specific expression of PSA has been the key
factor that led to the utility of PSA in early detection of CaP as
well as in the follow-up of disease progression and minimal
residual disease after treatment (2). The discovery of genes like
PSMA (4), NKX3.1 (6), prostate stem-cell antigen (8), prostase
(9), DD3 (11), and now PCGEM1 provides a panel of prostatespecific genes that may have potential to improve the diagnostic兾prognostic capability of PSA. Relative expression levels of
PSA, PSMA, NKX3.1, and PCGEM1 suggest a low abundance of
PCGEM1 in normal prostate tissue. It is also important to note
that PCGEM1 exhibited a significant tumor-associated overexpression in analysis of matched tumor兾normal specimens of
individual patients in both RT-PCR (56%) and in situ hybridization (84%) assays. The RT-PCR and in situ hybridization
assays were performed with two different sets of patient samples.
Apparent differences in the observed frequency of tumorassociated PCGEM1 expression by RT-PCR and in situ hybridization-based assays most likely reflect heterogeneity of PCGEM1 expression in normal and tumor-prostate tissues. Because
tissue specimens were microdissected for RT-PCR assay, anticipated variations in tissue sampling might result in the underestimation of PCGEM1 expression. To address these issues,
future experiments will analyze a larger cohort of specimens by
using real time PCR and in situ hybridization assays. The
12220 兩 www.pnas.org
Fig. 7. Androgen regulation of PCGEM1. LNCaP cells were cultured in RPMI
medium 1640 containing 10% (vol兾vol) charcoal-stripped FBS for 4 days and were
followed by treatment with synthetic androgen (R1881 for 24 h at 10 nanomolar
concentrations). Poly(A)⫹ RNA from treated and untreated cells was analyzed for
PCGEM1 expression by Northern blot hybridization as described for Fig. 4.
Srikantan et al.
variations in the levels of PCGEM1 expression in normal prostate tissue of different individuals need to be examined further.
Focal expression of PCGEM1 was also noted in both normal and
tumor cells. Comparison of the benign and tumor glands of the
same patient generally showed weak or no PCGEM1 expression
in normal cells and increased expression in tumor cells. PCGEM1 overexpression was observed in prostatic intraepithelial
neoplasia foci of some specimens. Although, in our limited
analysis, poorly differentiated cancer cells tended to show relatively higher expression of PCGEM1, a larger study is warranted
to correlate PCGEM1 expression with the clinicopathologic
features.
Prostate-tissue-specific expression of the PSA and probasin
genes have provided further opportunities to use promoters of
these genes for the prostate-tissue-specific targeting of genes in
transgenic mice models of CaP. Moreover, PSA promoter-driven
anticancer genes are being evaluated for their selective effects on
CaP cells. Finally, prostate-specific expression of PSA, PSMA,
and HK2 genes is being explored for the detection of CaP
micrometastasis (19). Therefore, studies of PCGEM1 by similar
approaches hold promising applications in both translational and
basic research areas of CaP.
The most intriguing aspect of PCGEM1 characterization has
been its lack of protein-coding capacity. Although we have not
completely ruled out whether PCGEM1 codes for a short
unstable peptide, careful sequencing of PCGEM1 cDNA (Fig. 1)
and genomic clones (Z.Z., unpublished work), computational
analysis of PCGEM1 sequence (Fig. 2 A and B), and in vitro
transcription兾translation experiments (data not shown) strongly
suggest a noncoding nature of PCGEM1. It is interesting to note
that an emerging group of mRNA-like noncoding RNAs are
being discovered whose function and mechanisms of action
remain poorly understood (20). Such RNA molecules have also
been termed as ‘‘RNA riboregulators’’ because of their function(s) in development, differentiation, DNA damage, heatshock responses, and tumorigenesis (21–24). In the context of
tumorigenesis, the H19, His-1, and Bic genes code for functional
noncoding mRNAs (24). In addition, a recently reported CaPassociated gene, DD3, also seems to exhibit a tissue-specific
noncoding mRNA (11). In this regard, it is important to point out
that PCGEM1 and DD3 may represent a distinct class of
prostate-specific genes. The recent discovery of a steroid receptor coactivator as an mRNA, lacking protein-coding capacity,
further emphasizes the role of RNA riboregulators in critical
biochemical function(s) (25). Our recent preliminary results
showed that PCGEM1 expression in NIH 3T3 cells caused a
significant increase in size of colonies in colony-forming assay,
suggesting that PCGEM1 cDNA confers cell proliferation
and兾or cell survival function(s) (V.S., unpublished work). Elevated expression of PCGEM1 in CaP cells may represent a gain
in function favoring tumor-cell proliferation兾survival. On the
basis of our first characterization of the PCGEM1 gene, we
propose that PCGEM1 belongs to a distinct class of prostatetissue-specific genes with potential functions in prostate cell
biology and the tumorigenesis of the prostate gland.
1. Landis, S. H., Murray, T., Bolden, S. & Wingo, P. A. (1999) Ca Cancer J. Clin.
49, 8–31.
2. Garnick, M. B. & Fair, W. R. (1998) Sci. Am. 279 (6), 74–83.
3. Augustus, M., Moul, J. W. & Srivastava, S. (1999) in Molecular Pathology of
Early Cancer, eds. Srivastava, S., Henson, D. E. & Gazden, A. (IOS, Amsterdam), pp. 321–340.
4. Fair, W. R., Israeli, R. S. & Heston, W. D. (1997) Prostate 32, 140–148.
5. Murphy, G. P., Tjoa, B. A., Simmons, S. J., Jarisch, J., Bowes, V. A., Ragde,
H., Rogers, M., Elgamal, A., Kenny, G. M., et al. (1999) Prostate 38, 73–78.
6. He, W. W., Sciavolino, P. J., Wing, J., Augustus, M., Hudson, P., Meissner,
P. S., Curtis, R. T., Schell, B. K., Bostwick, D. G., et al. (1997) Genomics 43,
69 –77.
7. Bhatia-Gaur, R., Donjacour, A. A., Sciavolino, P. J., Kim, M., Desai, N., Young,
P., Norton, C. R., Gridley, T., Cardiff, R. D., Cunha, G. R., et al. (1999) Genes
Dev. 13, 966–977.
8. Reiter, R. E., Gu, Z., Watabe, T., Thomas, G., Szigeti, K., Davis, E., Wahl, M.,
Nisitani, S., Yarnashiro, L., LeBeau, M. M., et al. (1998) Proc. Natl. Acad. Sci.
USA 95, 1735–1740.
9. Nelson, P. S., Gan, L., Ferguson, C., Moss, P., Gelinas, R., Hood, L. & Wang,
K. (1999) Proc. Natl. Acad. Sci. USA 96, 3114–3119.
10. Lin, B., Ferguson, C., White, J. T., Wang, S., Vessella, R., True, L. D., Hood,
L. & Nelson, P. S. (1999) Cancer Res. 59, 4180–4184.
11. Bussemakers, M. J. H., Van Bokhoven, A., Verhaegh, G. W., Smit, F. P.,
Karthaus, H. F., Schalken, J. A., Debruyne, F. M., Ru, N. & Isaacs, W. B. (1999)
Cancer Res. 59, 5975–5979.
12. Srikantan, V., Sesterhenn, I. A., Davis, L., Hankins, G. R., Avallone, F. A.,
Livezey, J. R., Connelly, R., Mostofi, F. K., McLeod, D. G., Moul, J. W., et al.
(1999) Int. J. Cancer. 84, 331–335.
Gingrich, J. R., Tucker, J. A., Walther, P. J., Day, J. W., Poulton, S. H. M. &
Webb, K. S. (1991) J. Urol. 146, 915–919.
Liang, P. & Pardee, A. B. (1992) Science 257, 967–971.
Heselmeyer, K., Macville, M., Schrock, E., Blegen, H., Hellstrom, A. C., Shah,
K., Auer, G. & Ried, T. (1997) Genes Chromosomes Cancer 4, 233–240.
Mitelman, F. (1995) in An International System for Human Cytogenetics
Nomenclature, ed. Mitelman, F. (Karger, Basel).
Gaddipati, J. P., McLeod, D. G., Sesterhenn, I. A., Hussussian, C. J., Tong,
Y. A., Seth, P., Dracopoli, N. C., Moul, J. W. & Srivastava, S. (1997) Prostate
30, 188–194.
Wilkinson, D. & Green, J. (1990) in Postimplantation Mammalian Embryos,
eds. Copp, A. J. & Cokroft, D. L. (Oxford Univ. Press, London), pp. 155–171.
de la Taille, A., Olsson, C. A. & Katz, A. E. (1999) Oncology 13, 187–194.
Erdmann, V. A., Szymanski, M., Hochberg, A., de Groot, N. & Barciszewski,
J. (1999) Nucleic Acids Research 27, 192–195.
Crespi, M. D., Jurkevitch, E., Poiret, M., d’Aubenton-Carafa, Y., Petrovics, G.,
Kondorosi, E. & Kondorosi, A. (1994) EMBO J. 13, 5099–5112.
Velleca, M. A., Wallace, M. C. & Merlie, J. P. (1994) Mol. Cell. Biol. 14,
7095–7104.
Takeda, K., Ichijo, H., Fujii, M., Mochida, Y., Saitoh, M., Nishitoh, H.,
Sampath, T. K. & Miyazono, K. (1998) J. Biol. Chem. 273, 17079–17085.
Askew, D. S. & Xu, F. (1999) Histol. Histopathol. 14, 235–241.
Lanz, R. B., McKenna, N. J., Onate, S. A., Albrecht, U., Wong, J., Tsai, S. Y.,
Tsai, M. J. & O’Mally, B. W. (1999) Cell 97, 17–27.
Xu, L., Srikantan, V., Sesterhenn, I. A., Augustus, M., Dean, R., Moul, J. W.,
Carter, K. C. & Srivastava, S. (2000) J. Urol. 63, 972–979.
We thank Mrs. Anita Roundtree for preparing this manuscript, Mr. Naga
Shanumgam for in vitro transcription兾translations, and Drs. Harvey B.
Pollard and Zoltan Szallasi for insightful discussions related to the
structure of PCGEM1 cDNA. This research was supported in part by a
grant from the Center for Prostate Disease Research, a program of the
Henry M. Jackson Foundation for the Advancement of Military Medicine (Rockville, MD), which is funded by the U.S. Army Medical
Research and Materiel Command.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
MEDICAL SCIENCES
26.
Srikantan et al.
PNAS 兩 October 24, 2000 兩 vol. 97 兩 no. 22 兩 12221